Hydrosilylation of a Dinuclear Tantalum Dinitrogen Complex

Feb 21, 2003 - bridging units can be considered as bridging silylimides bracketing a Ta(IV)-Ta(IV) metal-metal bond, thus accounting for the observed ...
2 downloads 0 Views 62KB Size
Published on Web 02/21/2003

Hydrosilylation of a Dinuclear Tantalum Dinitrogen Complex: Cleavage of N2 and Functionalization of Both Nitrogen Atoms Michael D. Fryzuk,* Bruce A. MacKay, and Brian O. Patrick Department of Chemistry, UniVersity of British Columbia, 2036 Main Mall, VancouVer, BC, Canada V6T 1Z1 Received January 23, 2003 ; E-mail: [email protected]

Research into the activation of molecular nitrogen by metal complexes has reached a high level of sophistication.1 While early work concentrated on the discovery of new examples of N2 coordinated to metals in a variety of bonding modes,2 it is clear that further advancement in this area will require systems that show enhanced reactivity patterns. In particular, discovery of metal dinitrogen complexes that undergo transformations resulting in the cleavage and/or functionalization3 of the coordinated N2 unit is of interest especially if the conditions are mild and potentially applicable to catalytic deployment.4 We recently reported the hydroboration of the tantalum dinitrogen complex ([NPN]Ta)2(µ-H)2(µ-η1:η2-N2), 1 (where [NPN] ) (PhNSiMe2CH2)2PPh), in which addition of the B-H bond of 9-BBN across a metal-dinitrogen π bond gave a new N-B bond and a terminal tantalum hydride.5 A key finding was that this reaction triggered a cascade of events including N-N bond cleavage. Unfortunately, an ancillary [NPN] ligand was irreversibly rearranged via silyl migration to a metal nitrido ligand, limiting the usefulness of this process. However, the success of this reaction provided the incentive to examine other simple hydride reagents that might add E-H bonds to 1 without the unattractive features of the hydroboration chemistry. We have previously reported hydrosilylation of dinitrogen complexes of zirconium,6 and thus exploring addition reactions with primary silanes seemed attractive. Herein we report preliminary results that extend this reactivity pattern for the N2 ligand and establish new possibilities for cleavage and complete functionalization of coordinated dinitrogen without the involvement of the ancillary ligands. The reaction between 2 equiv of butylsilane (BuSiH3) and the ditantalum dinitrogen complex 1 proceeds cleanly over 36 h to produce 2 in 63% yield (eq 1). The NMR spectral data of 2, particularly the presence of a single [NPN] silyl resonance and a single butylsilyl resonance in the 29Si{1H} NMR spectrum, are consistent with C2h symmetry in solution.7 No 1H NMR resonances typical of hydrides, either bridging or terminal, are observed.

The solid-state molecular structure of 2, determined from a twinned crystal, is shown in Figure 1.8 It is evident that the dinitrogen fragment has been cleaved and that both dinitrogenderived N atoms have been functionalized in the same manner despite their asymmetric activation in parent complex 1. These bridging units can be considered as bridging silylimides bracketing a Ta(IV)-Ta(IV) metal-metal bond, thus accounting for the observed diamagnetism. Labeling studies using the enriched dini3234

9

J. AM. CHEM. SOC. 2003, 125, 3234-3235

Figure 1. ORTEP drawing of 2 (ellipsoids at 50% probability, silyl methyl and phenyl carbons other than ipso omitted for clarity). Selected bond lengths (Å) and angles (deg): Ta1-N1 1.997(9), Ta1-N2 2.007(9), Ta2N1 2.038(10), Ta2-N2 1.990(9), N1‚‚‚N2 2.935(4), N1-Si1 1.729(10), N2-Si2 1.739(10), Ta1-Ta2 2.6677(6), Ta1-P1 2.735(3), Ta1-N3 2.070(9), Ta1-N4 2.045(9), Ta2-P2 2.687(3), Ta2-N5 2.077(9), Ta2-N6 2.069(8), Ta1-N1-Ta2 82.8(4), N1-Ta2-N2 93.5(4), Ta2-N2-Ta1 83.7(3), N2-Ta1-N1 94.3(4), Ta1-N1-Si1 143.2(6), Ta2-N2-Si2 134.8(5), Ta1-N2-Ta2-N1 155.8(5).

trogen complex 15N2-1 clearly indicate that the coordinated dinitrogen is the source of the N-atoms of the bridging silylimide units of 2. Other silanes such as PhSiH3 and Ph2SiH2 also add across the Ta2N2 moiety with similar outcomes as will be reported in due course. Insight into the intervening processes in this reaction was obtained by periodically observing the 31P{1H} NMR spectra of the reaction of butylsilane and 1. The first step is an addition reaction analogous to the hydroboration sequence5 that gives a new N-Si bond and new terminal hydride; the characteristic 31P resonances of 1 convert cleanly to those of complex 3 within 8 h at room temperature. This species has been isolated in excellent yield as crystals from a cooled solution of pentane; its solid-state molecular structure is shown in Figure 2.9 Complex 3 maps closely onto the N-boryl congener, but the butylsilyl group is clearly not coplanar with the Ta2N2 core. The Ta-N-Si angle is 133.6(5)° versus 163.1(10)° in the N-boryl analogue, and this may indicate a lack of N-Si π overlap and thus a change in hybridization at N1 between the two otherwise very similar complexes. The hydride ligands shown in Figure 2 were not located in the diffraction experiment but were successfully modeled using XHYDEX10 and are strongly implied by the 1H NMR spectrum of 3. Complex 3 is thermally unstable and undergoes spontaneous elimination of H2, cleanly converting to 4 over 8 h at room temperature; this process occurs both in solution and in the solid state. This new species exhibits Cs symmetry in solution in its 1H NMR spectrum; a resonance typical of a terminal tantalum hydride is observed at δ 17.4 ppm, but no resonances suggestive of bridging 10.1021/ja034303f CCC: $25.00 © 2003 American Chemical Society

COMMUNICATIONS Scheme 1

Figure 2. ORTEP drawing of 3 (ellipsoids at 50% probability, silyl methyl and phenyl carbons other than ipso omitted for clarity). Hydrides were placed using XHYDEX. Selected bond lengths (Å) and angles (deg): N1-N2 1.363(13), Ta1-N1 2.181(10), Ta1-N2 2.179(10), Ta2-N2 1.835(10), N1-Si1 1.732(11), Ta1‚‚‚Ta2 2.8430(9), Ta1-P1 2.583(3), Ta1-N3 2.028(10), Ta1-N4 1.236(11), Ta2-P2 2.632(3), Ta2-N5 2.044(8), Ta2-N6 2.075(7), Ta1-N1-Si1 133.6(5), Ta1-N1-N2 71.7(6), Ta1-N2-Ta2 89.7(4), N1-Ta1-P1 79.3(3), N1-Ta1-N3 159.5(4), N1-Ta1-N4 87.0(4), N2-Ta2-P2 166.6(3), N2-Ta2-N5 108.4(4), N2-Ta2-N6 105.8(4), N1-Ta1-Ta2-N2 1.0(5), Ta2-Ta1-N1-Si1 105.7(8), P1-Ta1Ta2-P2 -170.6(2).

hydrides or N-H species are present. We have previously proposed that loss of the bridging hydride ligands via reductive elimination as H2 provides the final two electrons required for overall reductive cleavage of the N-N bond. Evidence for this cleavage event comes from 15N{1H} NMR spectra of the isotopomers 15N2-3 and 15N2-4 prepared using enriched dinitrogen complex 15N2-1. Scalar coupling of 1JNN ) 16.6 Hz between the two resonances in 15N2-3 presumably results from the intact N-N bond. However, the spectrum of 15N2-4 shows no N-N scalar coupling and therefore indicates that N-N bond scission has occurred.11 Disilylimide 2 therefore appears to derive from intermediate 4 via a second addition of BuSiH3 across the Ta-NdTa unit,12 a dinuclear reductive elimination of the hydride ligands as H2 completes the process (Scheme 1).13 These results suggest three general conclusions regarding the addition of E-H bonds across the dinitrogen unit of 1. First, the initial reaction involves a common intermediate formed by the addition of 1 equiv of E-H across the π-bond of the ditantalum dinitrogen moiety. Second, this addition triggers loss of the bridging hydride ligands as H2, which provides the two electrons necessary to completely cleave the N-N bond. Third, the reducing power of this system is “built-in” via the reductive elimination of H2, which can occur repeatedly. As 1 is itself derived from a tantalum(IV) hydride dimer that forms via reductive elimination,14 this finding may bear strongly on the development of catalytic processes for the incorporation of dinitrogen-derived N atoms into other substrates. Such studies are in progress as are investigations into reactions of other organometallic E-H reagents with 1.

Acknowledgment. The authors thank NSERC of Canada for funding (Discovery Grant to M.D.F., PGS-A/B to B.A.M.), Mr. Christopher D. Carmichael for assistance with the crystallography, Mr. Marshall Lapawa for mass spectrometry, and Mr. Peter Borda and Mr. Minaz Lakha for elemental analyses. Supporting Information Available: X-ray and complete experimental details for 2, 3, 4, and isotopomers (PDF). Structural information for 2 and 4 (CIF). This material is available free of charge via the Internet at http://pubs.acs.org. References (1) (a) Fryzuk, M. D.; Johnson, S. A. Coord. Chem. ReV. 2000, 200-202, 379. (b) Hidai, M. Coord. Chem. ReV. 1999, 185-186, 99. (c) Hidai, M.; Mizobe, Y. Chem. ReV. 1995, 95, 1115. (d) Bazhenova, T. A.; Shilov, A. E. Coord. Chem. ReV. 1995, 144, 69. (e) Richards, R. L. Coord. Chem. ReV. 1996, 154, 83-97. (f) Richards, R. L. Pure Appl. Chem. 1996, 68, 1521. (g) Pickett, C. J. J. Biol. Inorg. Chem. 1996, 1, 601. (2) Chatt, J.; Dilworth, J. R.; Richards, R. L. Chem. ReV. 1978, 78, 589. (3) (a) Caselli, A.; Solari, E.; Scopelliti, R.; Floriani, C.; Re, N.; Rizzoli, C.; Chiesi-Villa, A. J. Am. Chem. Soc. 2000, 122, 3652-3670. (b) Cummins, C. C. Chem. Commun. 1998, 17, 1777-1786. (c) Dube, T.; Conochi, S.; Gambarotta, S.; Yap, G. P. A.; Vasapollo, G. Angew. Chem., Int. Ed. 1999, 38, 3657-3659. (d) Bates, V. M. E.; Clentsmith, G. K. B.; Cloke, F. G. N.; Green, J. C.; Huw, D. L. Chem. Commun. 2000, 11, 927-928. (4) (a) Yandulov, D. V.; Schrock, R. R. J. Am. Chem. Soc. 2002, 124, 62526253. (b) Floriani, C. In Modern Coordination Chemistry - the Legacy of Joseph Chatt; Leigh, G. J., Winterton, N., Eds.; RSC: Cambridge, U.K., 2002; pp 198-207. (5) Fryzuk, M. D.; MacKay, B. A.; Johnson, S. A.; Patrick, B. O. Angew. Chem., Int. Ed. 2002, 41, 3709-3712. (6) Fryzuk, M. D.; Love, J. B.; Rettig, S. J.; Young, V. G. Science 1997, 275, 1445-1447. (7) 29Si NMR spectral data for 2: (C6D6, 30 °C, 79.5 MHz) δ 8.70, (d, 2JPSi ) 7.0 Hz, NSiMe2), -1.74 (dd, 2JPSi ) 4.1 Hz, dd, 2JPSi ) 3.9 Hz, BuSiµN). (8) Crystallographic data for 2: C56H84N6P2Si6Ta2, triclinic, space group P-1, a ) 11.843(1) Å, b ) 11.853(1) Å, c ) 23.382(2) Å, R ) 86.68(1)°, β ) 76.24(1)°, γ ) 79.36(1)°, Z ) 4, two-component twin, related by 44° rotation normal to (-0.1, -1.0, -1.2), R1, 0.097 (I > 2.00σ(I)), wR2 0.211. (9) Crystallographic data for 3: C52H76N6P2Si5Ta2, orthorhombic, space group Pna21, a ) 27.600(6) Å, b ) 17.262(4) Å, c ) 14.701(3) Å, Z ) 4, R1, 0.0566 (I > 2.00σ(I)), wR2 0.1353. (10) Orpen, A. G. J. Chem. Soc., Dalton Trans. 1980, 2509. 15 (11) N NMR spectral data for 15N2-3: (C7D8, -60 °C, 40 MHz) δ -163.5 (d, 1JNN ) 16.6 Hz), -52.5 (dd, 2JPN ) 26.4 Hz, 1JNN ) 16.6 Hz). 15N NMR spectral data for 15N2-4: (C6D6, 30 °C, 40 MHz) δ 284.4 (b), -44.8 (d, 2JPN ) 18.7 Hz). (12) Gountchev, T. I.; Tilley, T. D. J. Am. Chem. Soc. 1997, 119, 1283112841. (13) de Wolf, J. M.; Blaauw, R.; Meetsma, A.; Teuben, J. H.; Gyepes, R.; Varga, V.; Mach, K.; Veldman, N.; Spek, A. L. Organometallics 1996, 15, 4977-4983. (14) Fryzuk, M. D.; Johnson, S. A.; Patrick, B. O.; Albinati, A.; Mason, S. A.; Koetzle, T. F. J. Am. Chem. Soc. 2001, 123, 3960-3973.

JA034303F

J. AM. CHEM. SOC.

9

VOL. 125, NO. 11, 2003 3235